You are on page 1of 8

High pressure antiferrodistortive phase transition in mixed crystals of EuTiO3 and

SrTiO3
,
Paraskevas Parisiades , Francesco Saltarelli, Efthymios Liarokapis, Jürgen Köhler, and Annette Bussmann-
Holder

Citation: AIP Advances 6, 065019 (2016); doi: 10.1063/1.4954819


View online: http://dx.doi.org/10.1063/1.4954819
View Table of Contents: http://aip.scitation.org/toc/adv/6/6
Published by the American Institute of Physics
AIP ADVANCES 6, 065019 (2016)

High pressure antiferrodistortive phase transition in mixed


crystals of EuTiO3 and SrTiO3
Paraskevas Parisiades,1,2,a Francesco Saltarelli,3 Efthymios Liarokapis,4
Jürgen Köhler,5 and Annette Bussmann-Holder5
1
Laboratoire Matériaux et Phénoménes Quantiques (UMR 7162 CNRS), Université Paris
Diderot-Paris 7, Paris Cedex 13, France
2
ID27 Beamline, European Synchrotron Radiation Facility, 71 Avenue des Martyrs,
38000 Grenoble, France
3
Sapienza University of Rome, Physics Dept., Piazzale Aldo Moro 5, 00185 Roma, Italy
4
Department of Physics, National Technical University of Athens, GR-15780 Athens, Greece
5
Max-Planck-Institute for Solid State Research, Heisenbergstrasse 1,
D-70569 Stuttgart, Germany
(Received 3 May 2016; accepted 14 June 2016; published online 21 June 2016)

We report a detailed high pressure study on Eu1-xSrxTiO3 polycrystalline samples


using synchrotron x-ray diffraction. We have observed a second-order antiferrodis-
tortive phase transition for all doping levels which corresponds to the transition
that has been previously explored as a function of temperature. The analysis of
the compression mechanism by calculating the lattice parameters, spontaneous
strains and tilt angles of the TiO6 octahedra leads to a high pressure phase dia-
gram for Eu1-xSrxTiO3. C 2016 Author(s). All article content, except where oth-
erwise noted, is licensed under a Creative Commons Attribution (CC BY) license
(http://creativecommons.org/licenses/by/4.0/). [http://dx.doi.org/10.1063/1.4954819]

I. INTRODUCTION
Perovskite materials of the ABO3 general formula are often sensitive to structural changes with
temperature and pressure because of the tilting of BO6 octahedra. SrTiO3 (STO) is one of the most
studied perovskites that undergoes such an antiferrodistortive structural phase transition at 105 K1
or 9.5 GPa2 from cubic Pm-3m to tetragonal I4/mcm. EuTiO3 (ETO) on the other hand has attracted
a lot of interest only recently, mainly due to the recent discovery of strong magneto-electric coupl-
ing at low temperature in this compound.3 The comparison of STO and ETO is very important since
these compounds share many common properties: they crystallize in the cubic ABO3 perovskite
structure, they share similar lattice constants and ionic radii for Eu2+ and Sr2+, and the Ti4+ ion is
in the d0 configuration. The latter is a crucial condition for ferroelectricity, and both compounds
show the tendency to a ferroelectric instability as evidenced by a transverse optic mode softening.4–6
Their main difference lies in the magnetic nature of 4f7 Eu2+ ion with strongly localized spins
of S=7/2, which is responsible for the appearance of a G-type antiferromagnetism (AFM) below
TN=5.5 K.3,7,8
It has been recently shown by several different experimental techniques9–15 that the same
structural phase transition as observed in STO at TS = 105K, is also present in ETO, but at much
higher temperature, namely TS = 220 – 285 K. We also showed recently that, as in the case of STO,
there is a pressure-induced transition in ETO starting roughly at ∼2.7 GPa.16 The phase transition
of ETO is not very well understood, and depends on the synthesis conditions,5 the presence of
mixed valence states for Eu,11 or oxygen vacancies.11 In any case, the large difference in the critical
temperature/pressure for the phase transitions in STO and ETO is unexpected and it is possible that
spin-phonon coupling effects govern the structural instability of the latter.3,9

a Electronic mail: paraskevas.parisiadis@univ-paris-diderot.fr.

2158-3226/2016/6(6)/065019/7 6, 065019-1 © Author(s) 2016.


065019-2 Parisiades et al. AIP Advances 6, 065019 (2016)

The peculiar property of ETO to show “hidden” magnetism far above the AFM state together
with the suppression of AFM order by Sr doping17 is expected to reveal novel states when being
driven into the vicinity of a quantum critical point. The close analogy of ETO with STO suggests
that the multiple application properties of STO should also be present in ETO based materials where
the additional spin degree of freedom offers new perspectives and functionalities as compared to
STO. In this work we fill in the gap of information between STO and ETO by carrying out a thor-
ough synchrotron x-ray diffraction (XRD) study under hydrostatic pressure for the solid solution
Eu1-xSrxTiO3. Our data clearly show the antiferrodistortive phase transition as a function of x and
establish a new pressure phase diagram of Eu1-xSrxTiO3.

II. EXPERIMENTAL
Polycrystalline samples of Eu1-xSrxTiO3 (x=0.25, 0.50, 0.75 and 0.97), with crystallite sizes
2-5 µm have been prepared as described in Ref 9. The synchrotron x-ray diffraction experiments
were performed on the ID27 beamline at the ESRF, using a monochromatic wavelength of 0.3738 Å
(33.168 keV). The beam size was focused to a spot of about 3x3 µm2. The samples were loaded in
Le Toullec type membrane diamond anvil cells (DACs) with diamond culets of 300 µm in diameter.
The rhenium gaskets were pre-indented to a thickness of 50 µm, while the sample chamber was
created by drilling a 150 µm hole in the gasket with a Nd:YAG pulsed laser. Helium gas has been
used as a pressure transmitting medium (PTM), thus minimizing the non-hydrostatic stresses that
could modify the onset or the sequence of the phase transitions in perovskites. The pressure has
been measured using the fluorescence lines of ruby. The data were collected with a Perkin Elmer
flat panel detector and the acquired images have been integrated using the Dioptas software.18
The diffractograms were indexed using DICVOL91 software.19 The Lebail refinements have been
carried out with the Fullprof software package.20

III. RESULTS AND DISCUSSION


The Eu1-xSrxTiO3 samples have been measured at room temperature up to a maximum pressure
of 48 GPa. The phase transition from cubic to tetragonal in perovskites can be followed by x-ray
diffraction either by the splitting of Bragg peaks or by the appearance of superstructure reflections.
In Fig. 1 we display the evolution of the (200) peak in the diffraction patterns of all four samples
for some selected pressures. The data clearly show the splitting of the (200) reflection to (220) and
(004), that accounts for the tetragonal distortion in Eu1-xSrxTiO3. The same phase transition has
been detected under pressure for pure EuTiO316 and SrTiO3.2 As in the case of EuTiO3, the tran-
sition is fully reversible as can be seen in the decompression data in the supplementary material.21
The phase transition is generated by the tilting of the TiO6 octahedra with a tilting scheme a0a0c- in
the Glazer notation,22 i.e. an out-of-phase rotation of the c axis with no rotation of the a and b axes.
In order to accurately define the onset of the pressure-induced transition, we plot the normal-
ized linewidth of the (200) reflection with pressure, as shown in Fig. 2(a). The same method has
also been followed in the case of EuTiO3,16 since it minimizes any resolution issues that could delay
the emergence of the transition. Fitting the (200) peak with a single profile function will result in
an abrupt increase in the FWHM at the pressure where the tetragonal phase appears, and the (200)
reflection becomes a doublet. The kinks observed in the FWHM plots of Fig. 2 clearly demonstrate
that the critical transition pressure Pc increases for increasing Sr content in Eu1-xSrxTiO3. In a
high pressure diffraction experiment peak broadening can occur due to the hardening of PTM or
the squeezing of the sample between the diamonds (“bridging”). Both of these possibilities are
excluded in our experiments: First of all, He used in this study is the best quasi-hydrostatic PTM
that remains in the liquid phase liquid until ∼12 GPa. Second, the gasket thickness before and after
the experiment was far superior than the initial thickness of the sample, meaning that bridging has
been prevented. Moreover, the transition is found to be reversible in our experiments without any
significant hysteresis, confirming the absence of peak broadening due to non-hydrostatic stresses.
A second criterion in the definition of the onset for the phase transition has been the Lebail
065019-3 Parisiades et al. AIP Advances 6, 065019 (2016)

FIG. 1. Evolution of the (200) diffraction peak of Eu1-xSrxTiO3 with pressure for a) x=0.25, b) x=0.50, c) x=0.75 and
d) x=0.97.

refinements, which are much better fitted with the tetragonal pattern I4/mcm beyond the critical
pressures. Refinements of the x-ray data using both the cubic and tetragonal models are shown in
the supplementary material.21
The normalized (pseudo-cubic) cell volumes with respect to pressure are given in Fig. 2(b). The
pseudo-cubic plots of the lattice constants vs pressure for all samples are shown in the supplemen-
tary material.21 Using an F-f plot (see supplementary material21), our data in the tetragonal phase
can be represented with a straight horizontal line, indicating that a 2nd order Birch-Murnaghan
(BM2) equation of state (EoS) is sufficient to reproduce the data. We obtain the bulk moduli for
both cubic and tetragonal phases that are summarized in Table I together with the critical pressures
Pc of the transition for each of the four investigated samples. The BM2 EoS yields bulk moduli
between 168-175 GPa for the cubic phase and 177-187 GPa for the tetragonal phase, values that are
comparable with previous experimental and theoretical studies of EuTiO3 or SrTiO3.2,16,23–25 The
higher values for the tetragonal phase are reasonable since the material becomes less compressible
at higher pressures. No discontinuities have been observed in the pseudo-cubic volume evolution
with pressure, indicating that the nature of the transition is of second order. It is important to
emphasize however that despite the apparently second order nature of the phase transition in our
work, recent thermal expansion measurements support the first order character of it.15
The tetragonal distortion in perovskites can be quantified by the use of symmetry-adapted
spontaneous strains, ea = (2e1 + e3) (volume strain) and et = 2(e3 − e2) (tetragonal strain), where
e1 and e2 are the strain components calculated by using the pseudo-cubic lattice constants apc
and cpc: e1 = e2 = (apc − a0)/a0 and e3 = (cpc − a0)/a0. In Fig. 3(a) we present the evolution of the
spontaneous strains with pressure for x=0.25. The volume strain is very small in all cases and can
be considered negligible within error bars for all studied pressures, meaning that the elongation
of the octahedra along the c axis compensates almost completely the volume reduction caused by
the tilts. The tilt angles can be obtained by the exact atomic positions since they depend on the
displacements of O atoms. However, in our data, the polycrystalline nature of the samples made it
difficult to obtain Rietveld refinements with satisfying uncertainties at high pressures. Alternatively,
065019-4 Parisiades et al. AIP Advances 6, 065019 (2016)

FIG. 2. a. Normalized FWHM of the (200) diffraction peak, indicating the onset of the phase transition in Eu1-xSrxTiO3.
b. Pseudo-cubic volume vs pressure at room temperature for Eu1-xSrxTiO3. The open symbols correspond to the cubic phase
and the closed ones to the tetragonal phase. The dashed-dotted lines represent the BM2 fits. The errors bars are smaller than
the symbols.

the tilt angles can be calculated by the lattice constants by assuming the octahedra to remain undis-
torted (i.e. octahedral that maintain their corner-sharing connectivity).
√ For a tetragonal lattice the
octahedral tilting can be expressed by the relation φ = arctan( 2a/c). The tilt angle, which is the
order parameter of the observed structural transition, clearly shows that the tetragonal distortion is

TABLE I. Critical transition pressures and bulk moduli for Eu1-xSrxTiO3.

x Pc (GPa) K0 Pm-3m (GPa) K0 I4/mcm (GPa)

0.25 3.7 168.7 (8) 177.8 (7)


0.50 4.6 172.5 (6) 181.0 (3)
0.75 5.5 175.4 (9) 187.8 (9)
0.97 7.6 174.3 (5) 183.0 (4)
065019-5 Parisiades et al. AIP Advances 6, 065019 (2016)

FIG. 3. a. Evolution of spontaneous strains with pressure for x=0.25. b. Evolution of the octahedral tilt angle with pressure
for x=0.25. The error bars are smaller than the symbols.

small for low pressures and strongly increases at higher pressures, reaching a maximum value of
∼12o for the maximum pressure (Fig. 3(b)). The data for the remaining samples are presented in the
supplementary material.21
From our data it is possible to construct a high pressure phase diagram of Eu1-xSrxTiO3, as
shown in Fig. 4. We add also for comparison the data points for pure EuTiO316 and SrTiO3.2 Our
data can be fitted with a straight line with a positive slope dPc /dx=5.2 GPa. Extrapolating to x=0
(pure EuTiO3) gives a transition pressure of Pc =2.2 GPa, which is close to the 2.7 Gpa of our
previous work.16 However, by extrapolating for x=1 we obtain Pc = 7.6 GPa for pure SrTiO3, which
is substantially lower than 9.6 GPa as reported by Guennou et al.2 The difference can most likely
be ascribed to the different PTM used in the two experimental works (He in our work and Ne
in Guennou et al.2). Oxygen deficiencies or oxygen excess accompanied by the presence of small
amounts of Eu3+ ions in the Eu1-xSrxTiO3 samples are additional error sources leading to slight
variation of the lattice parameters and thus different transition pressures.11 These problems have
not been encountered in pure SrTiO3. A previous work on the low temperature phase diagram of
Eu1-xSrxTiO3 using electron paramagnetic resonance (EPR) and resistivity measurements26 show a
non-linear behavior of TS with x, which is unexpected given the many structural similarities be-
tween the two end members, mainly the similar ionic radii between Sr and Eu. Since the non-linear
behavior cannot be due to a lattice mismatch, the authors concluded that there is a change in the
dynamics around x=0.75 that is dependent on the next-nearest neighbor spin-phonon interactions. If
065019-6 Parisiades et al. AIP Advances 6, 065019 (2016)

FIG. 4. Phase diagram of Eu1-xSrxTiO3. Our present work (circles) is compared with our previous work on pure EuTiO3
(rectangle) and the work of Guennou et al. (triangle) on pure SrTiO3.

this is the case, our x-ray diffraction measurements at room temperature are not sensitive to such an
effect since Bragg reflections are a lot stronger than any spin-related effects and therefore, a linear
behavior with pressure is observed, as in the case of pure EuTiO316 and SrTiO3.2

IV. CONCLUSION
To summarize, we have made a detailed high pressure synchrotron x-ray diffraction study of
Eu1-xSrxTiO3. We have observed the antiferrodistortive cubic-to-tetragonal phase transition in the
whole doping regime which is identical to the phase transition observed in both end members
either under high pressure or at low temperature and corresponds to the tilting of TiO6 octahedra.
No further phase transitions have been observed up to 48 GPa within the resolution limits of our
experiment. We conclude with a P-x structural phase diagram for Eu1-xSrxTiO3, which exhibits a
linear behavior with x.
1 G. Shirane and Y. Yamada, Phys. Rev. 177, 858 (1969).
2 M. Guennou, P. Bouvier, J. Kreisel, and D. Machon, Phys. Rev B 81, 054115 (2010).
3 T. Katsufuji and H. Takagi, Phys. Rev. B 64, 054415 (2001).
4 R. A. Cowley, W. J. L. Buyers, and G. Dolling, Solid State Commun. 7, 181 (1969).
5 V. Goian, S. Kamba, O. Pacherová, J. Drahokoupil, L. Palatinus, M. Dušek, J. Rohlíček, M. Savinov, F. Laufek, W. Schranz,

A. Fuith, M. Kachlík, K. Maca, A. Shkabko, L. Sagarna, A. Weidenkaff, and A. A. Belik, Phys. Rev. B 86, 054112 (2012).
6 S. Kamba, D. Nuzhnyy, P. Vaněk, M. Savinov, K. Knížek, Z. Shen, E. Santavá, K. Maca, M. Sadowski, and J. Petzelt,

Europhys. Lett. 80, 27002 (2007).


7 T. R. McGuire, M. W. Shafer, R. J. Joenk, H. A. Halperin, and S. Pickart, J. Appl. Phys. 37, 981 (1966).
8 V. Scagnoli, M. Allieta, H. Walker, M. Scavini, T. Katsufuji, L. Sagarna, O. Zaharko, and C. Mazzoli, Phys. Rev. B 86,

094432 (2012).
9 A. Bussmann-Holder, J. Köhler, R. K. Kremer, and J. M. Law, Phys. Rev. B 83, 212102 (2011).
10 M. Allieta, M. Scavini, L. J. Spalek, V. Scagnoli, H. Walker, C. Panagopoulos, S. S. Saxena, T. Katsufuji, and C. Mazzoli,

Phys. Rev. B 85, 184107 (2012).


11 B. J. Kennedy, G. Murphy, E. Reynolds, M. Avdeev, H. E. R. Brand, and T. Kolodiazhnyi, J. Phys.: Condens. Matter 26,

495901 (2014).
12 J. W. Kim, P. Thomson, S. Brown, P. S. Nomile, J. A. Schlueter, A. Shkabko, A. Weidenkaff, and P. J. Ryan, Phys. Rev. Lett.

110, 027201 (2013).


13 L. J. Spalek, S. S. Saxena, C. Panagopoulos, T. Katsufuji, J. A. Schiemer, and M. A. Carpenter, Phys. Rev. B 90, 054119

(2014).
14 D. S. Ellis, H. Uchiyama, S. Tsutsui, K. Sugimoto, K. Kato, D. Ishigawa, and A. Q. R. Baron, Phys. Rev. B 86, 220301(R)

(2012).
15 P. G. Reuvekamp, R. K. Kremer, J. Köhler, and A. Bussmann-Holder, Phys. Rev. B 90, 104105 (2014).
16 P. Parisiades, E. Liarokapis, J. Köhler, A. Bussmann-Holder, and M. Mezouar, Phys. Rev. B 92, 064105 (2015).
17 A. Bussmann-Holder and J. Köhler, J. Phys. Chem. Solids 84, 2 (2015).
065019-7 Parisiades et al. AIP Advances 6, 065019 (2016)

18 G. Ashiotis, A. Deschildre, Z. Nawaz, J. P. Wright, D. Karkoulis, F. E. Picca, and J. Kieffer, J. Appl. Cryst. 48, 510 (2015).
19 A. Boultif and D. Louer, J. Appl. Cryst. 24, 987 (1991).
20 J. Rodriguez-Carvajal, Physica B 192, 55 (1993).
21 See supplementary material at http://dx.doi.org/10.1063/1.4954819 for decompression data, examples of Lebail refine-

ments, evolution of lattice constants with pressure, F-f plots, spontaneous strains and tilt angles.
22 A. M. Glazer, Acta Cryst. B 28, 3384 (1972).
23 T. Ishidate, S. Sasaki, and K. Inoue, High Press. Res. 1, 53 (1988).
24 L. R. Edwards and R. W. Lynch, J. Phys. Chem. Solids 31, 573 (1970).
25 A. G. Beattie and G. A. Samara, J. Appl. Phys. 42, 2376 (1971).
26 Z. Guguchia, A. Shengelaya, H. Keller, J. Köhler, and A. Bussmann-Holder, Phys. Rev. B 85, 134113 (2012).

You might also like